Fact-checked by Grok 2 weeks ago

Rossby number

The Rossby number (Ro) is a in that quantifies the ratio of inertial forces to Coriolis forces, playing a central role in analyzing rotating fluid flows such as those in the Earth's atmosphere and oceans. It is mathematically defined as Ro = \frac{U}{f L}, where U is a characteristic velocity scale, L is a characteristic length scale, and f = 2 \Omega \sin \phi is the Coriolis parameter, with \Omega denoting Earth's angular rotation rate and \phi the . The number is named after the Swedish-American Carl-Gustaf Arvid Rossby, who pioneered its application in geophysical contexts during the early . In geophysical fluid dynamics, a small Rossby number (typically Ro \ll 1) indicates that rotational effects dominate, leading to approximations like geostrophic balance where the counters pressure gradients, which is prevalent in large-scale atmospheric circulations and oceanic gyres. Conversely, a large Rossby number (Ro \gg 1) signifies negligible influence from , as seen in small-scale or high-speed flows like in non-rotating experiments or rapid convective motions. This parameter underpins key simplifications in modeling, such as the , which filter out high-frequency inertial oscillations to focus on balanced, slowly evolving dynamics essential for weather prediction and climate simulations. The Rossby number's utility extends to diverse applications, including the study of mid-latitude weather systems where Ro \approx 0.1 justifies rotational dominance, and equatorial dynamics where vanishing f requires modified approaches to capture rotational influences. It also informs the propagation and stability of phenomena like Rossby waves, which govern long-range teleconnections in global climate patterns, and aids in scaling analyses for planetary atmospheres beyond . By providing a measure of rotational constraint, the Rossby number remains a foundational tool for interpreting the interplay between local fluid motions and global planetary rotation.

Definition

Mathematical Expression

The Rossby number, denoted as \mathrm{Ro}, is defined by the formula \mathrm{Ro} = \frac{U}{f L}, where U represents the scale of the flow, L is the scale, and f is the Coriolis parameter. The Coriolis parameter f is expressed as f = 2 \Omega \sin \phi, in which \Omega is Earth's (approximately $7.292 \times 10^{-5} rad/s) and \phi is the . In this expression, U typically denotes the flow speed, such as or velocity (e.g., around 10 m/s), L indicates the horizontal scale of the system, such as the of a (e.g., $10^6 m for synoptic features), and f varies geographically, equaling zero at the (\phi = 0^\circ) and achieving its maximum of $2 \Omega at the poles (\phi = \pm 90^\circ). The combination yields a , as the units of U (m/s) are divided by those of f (s^{-1}) and L (m), resulting in cancellation.

Physical Interpretation

The Rossby number serves as a dimensionless that measures the relative importance of inertial (or advective) forces to Coriolis forces in rotating fluid systems, such as those in the atmosphere and . When the Rossby number is low, rotational effects dominate the , constraining motions to align with geostrophic principles; conversely, high values imply that local inertial accelerations overshadow the influence of planetary , permitting more isotropic or centrifugal-driven behaviors. This ratio provides a framework for classifying regimes and predicting whether must be explicitly accounted for in geophysical models. Threshold values of the Rossby number delineate distinct dynamical balances. For Ro ≪ 1, typically on the order of 0.1 or less, geostrophic balance prevails, wherein the counteracts the , resulting in straight-line flows perpendicular to the . In transitional regimes where Ro ≈ 1, both inertial and Coriolis terms contribute comparably, necessitating full momentum equations for accurate description. When Ro ≫ 1, often exceeding 10, cyclostrophic or inertial balances dominate, as seen in small-scale or low-latitude phenomena where the becomes negligible relative to centrifugal accelerations. These regimes manifest in natural systems with varying scales. Large-scale mid-latitude systems exhibit small Rossby numbers, emphasizing rotation's role in maintaining geostrophic and enabling phenomena like Rossby waves. In contrast, tropical cyclones display Rossby numbers around unity or higher in their cores, shifting toward cyclostrophic balance where radial pressure gradients balance centrifugal forces. Tornadoes represent extreme cases with Rossby numbers on the order of 10³, where inertial forces entirely eclipse Coriolis effects, allowing intense, rotationally unconstrained vortices. The applicability of the Rossby number rests on assumptions of primarily horizontal flow scales and steady or quasi-steady conditions, which align with large-scale geophysical contexts. However, significant vertical shears or rapid time variations can limit its direct use, often requiring extensions like quasi-geostrophic approximations to incorporate such complexities.

Theoretical Foundations

Derivation

The derivation of the Rossby number begins with the non-dimensionalization of the Navier-Stokes momentum equations for an incompressible fluid in a , where the plays a central role. The starting point is the horizontal momentum equation, simplified by neglecting viscosity (valid for high flows) and focusing on the balance between inertial acceleration, , and Coriolis effects: \frac{D\mathbf{u}}{Dt} = -\frac{1}{\rho}\nabla p + \mathbf{f} \times \mathbf{u}, where \mathbf{u} is the horizontal velocity vector, \frac{D}{Dt} = \frac{\partial}{\partial t} + (\mathbf{u} \cdot \nabla) is the material derivative representing inertial terms, p is pressure, \rho is density, and \mathbf{f} = f \hat{\mathbf{k}} with f = 2\Omega \sin\phi ( \Omega is Earth's angular velocity and \phi is latitude) is the Coriolis vector. This derivation assumes an f-plane approximation, where f is constant (neglecting latitudinal variations in the Coriolis ), and horizontal , meaning scales in the x- and y-directions are comparable. To non-dimensionalize, introduce characteristic scales: velocity magnitude U, horizontal length scale L, and advective T = L/U. Define dimensionless variables as \mathbf{u}' = \mathbf{u}/U, \mathbf{x}' = \mathbf{x}/L, t' = t/(L/U), and p' = p/(\rho U f L) ( to Coriolis and terms). Substituting these into the momentum equation yields the non-dimensional form after dropping primes for clarity: \text{Ro} \left( \frac{\partial \mathbf{u}}{\partial t} + \mathbf{u} \cdot \nabla \mathbf{u} \right) = -\nabla p + \hat{\mathbf{k}} \times \mathbf{u}, where the Rossby number emerges as \text{Ro} = U/(f L). The factor \text{Ro} multiplies the inertial (material derivative) term because the scaling of the advective acceleration is U^2/L, while the Coriolis acceleration scales as f U; their ratio is thus U/(f L), positioning \text{Ro} as the coefficient that balances these terms when the equation is normalized by the Coriolis scale. In the limit of small \text{Ro} \ll 1, the inertial term becomes negligible, leading to geostrophic balance where Coriolis and pressure gradient forces approximately cancel.

Relation to Other Dimensionless Numbers

The Rossby number (Ro) quantifies the ratio of inertial forces to Coriolis forces in rotating fluid flows, contrasting with the (Re = UL/ν), which measures the balance between inertial and viscous forces. In geophysical contexts, Ro is typically of order unity, indicating comparable inertial and rotational effects, while Re is very large due to low , leading to turbulent flows where viscosity plays a minor role except in boundary layers. The two numbers are interconnected through the relation Ro = Re × Ek, where Ek is the Ekman number, highlighting how rotation influences the transition from viscous to inertial dominance. The Ekman number (Ek = ν/(f L²)) represents the ratio of viscous to Coriolis forces and is generally very small (e.g., 10^{-10} to 10^{-5}) in geophysical flows, signifying that friction is negligible in the interior but critical near boundaries, forming Ekman layers of thickness δ ≈ √(2ν/f). Together, and Ek describe rotating-viscous balances: small Ek with moderate Ro implies near-geostrophic flow with thin frictional layers, as seen in large-scale atmospheric and oceanic circulations. In stratified or gravity-driven flows, Ro interacts with the (Ri = g Δρ H / (ρ₀ U²)), which assesses buoyancy versus shear instability, and the (Fr = U / √(g H)), which compares inertial to gravitational forces. These combinations enable classification of flow regimes; for instance, in ocean thermocline dynamics, O(1) Ro and Ri values indicate submesoscale processes where , , and inertia compete, driving frontogenesis and mixing. Composite regimes often require multiple numbers for full characterization. Geostrophic , prevalent in mid-latitude oceans and atmospheres, features low (<<1, rotation-dominated) and high Re (>>1, inertial), resulting in anisotropic cascades constrained by planetary conservation. Conversely, equatorial flows exhibit large (>>1) due to vanishing Coriolis f ≈ 0, allowing nearly non-rotating dynamics akin to shallow-water waves.
Dimensionless NumberDefinitionTypical Values in Geophysical FlowsRole in Multi-Parameter Regimes
Rossby (Ro)U / (f L)O(0.1–1)Rotation vs. inertia; low Ro with small Ek for geostrophy, high Ro with O(1) Ri/Fr for submesoscales.
Reynolds (Re)U L / ν>>1 (10^6–10^9)Inertia vs. viscosity; high Re with low Ro enables geostrophic turbulence.
Ekman (Ek)ν / (f L²)<<1 (10^{-10}–10^{-5})Viscosity vs. rotation; small Ek with moderate Ro defines interior inviscid balance with boundary layers.
Richardson (Ri)g Δρ H / (ρ₀ U²)O(0.1–10)Buoyancy vs. shear; O(1) Ri with O(1) Ro classifies unstable fronts in thermoclines.
Froude (Fr)U / √(g H)O(0.01–0.1)Inertia vs. gravity; low Fr with low Ro stabilizes stratified rotating flows.

Applications

Atmospheric Dynamics

In mid-latitude synoptic-scale flows, such as those associated with extratropical cyclones, the Rossby number typically ranges from 0.1 to 1, reflecting a balance where rotational effects are significant but inertial forces are comparable. This regime enables the , wherein the balances the , providing a foundational framework for understanding wind-pressure relationships in large-scale weather systems. In contrast, tropical systems like hurricanes exhibit higher Rossby numbers, often on the order of 10 to 100, due to the small Coriolis parameter f at low latitudes and large characteristic velocities U. Here, the reduced influence of Earth's rotation leads to gradient wind balance, where centrifugal forces play a prominent role alongside the Coriolis and pressure gradient forces, particularly in the inner core regions. Low Rossby number regimes are essential for the propagation of Rossby waves, which are planetary-scale undulations in the atmosphere driven by the variation of the Coriolis parameter with latitude. In these conditions, the deformation radius L_d = \frac{N H}{f}, where N is the Brunt-Väisälä frequency and H is the scale height, emerges from scale analysis as a critical length scale that delineates regions where rotational effects dominate wave dynamics. Waves with wavelengths much larger than L_d align with low Rossby number approximations, supporting the quasi-geostrophic framework for planetary wave evolution. Observational examples highlight the Rossby number's role in distinguishing flow regimes: in jet streams, where Ro < 1, planetary rotation is crucial for maintaining geostrophically balanced, large-scale zonal flows. Conversely, in thunderstorms, Ro > 1 due to smaller horizontal scales (around 10 km) and higher local velocities, allowing inertial and local convective dynamics to dominate over rotational influences. In , low Rossby numbers in mid-latitude synoptic flows justify the use of quasi-geostrophic models, which simplify the governing equations by assuming near-geostrophic balance and enable efficient forecasting of large-scale patterns. These models filter out high-frequency gravity waves, focusing computational resources on rotationally dominated evolutions critical for medium-range predictions.

Oceanic Circulation

In oceanic circulation, the Rossby number quantifies the relative importance of inertial forces to the Coriolis effect, influencing the dynamics of large-scale flows such as gyres and boundary currents. For flows with low Rossby numbers, geostrophic balance dominates, where pressure gradients equilibrate the Coriolis force, leading to nearly circular streamlines in the horizontal plane. Mesoscale eddies in subtropical gyres exhibit Rossby numbers typically ranging from 0.1 to 1, where nonlinear balances the , enabling the formation of coherent structures that transport properties across basins. These eddies often align with the Rossby radius of deformation, given by L_R = \frac{\sqrt{gH}}{f}, where g is , H is the characteristic water depth, and f is the Coriolis parameter; the Rossby number Ro = U / (f L) then determines the nonlinearity of eddies, with values near unity indicating a shift toward ageostrophic and enhanced variability. In western boundary currents such as the , Rossby numbers are higher, approximately 1 to 10, especially in regions of pronounced meanders, where inertial effects amplify instabilities and promote eddy shedding. This elevated Rossby number reflects the intense velocities and smaller horizontal scales, driving nonlinear interactions that intensify the meandering and contribute to cross-frontal exchanges. Near the equator, where f \approx 0, the Rossby number becomes very large (>100), minimizing the Coriolis influence and favoring inertia-gravity waves over rotational dynamics. altimetry observations highlight spatial variations in Rossby number across basins, which modulate eddy-driven mixing and the dispersion of tracers like heat and carbon.

Laboratory and Engineering Contexts

In laboratory settings, rotating experiments are widely used to simulate geostrophic flows dominated by the Coriolis effect, where the (Ro) is typically maintained at low values of approximately 0.01 to 0.1 to mimic large-scale geophysical phenomena such as cyclones and Rossby waves. These experiments often employ turntables or annuli to impose rotation, allowing researchers to observe balanced vortex dynamics and wave propagation under controlled conditions, with Ro = U/(fL) quantifying the relative weakness of inertial forces compared to Coriolis deflection, where U is , f is the Coriolis parameter (twice the rotation rate), and L is a horizontal length scale. For instance, in rotating annulus setups, Ro values as low as 7 × 10^{-4} to 0.02 enable the excitation of standing Rossby waves with mode numbers from 2 to 12, closely replicating theoretical predictions for gradients. Geophysical fluid dynamics laboratories further utilize these setups, such as parabolic mirrors or rotating tables, to achieve precise f-plane or β-plane approximations for studying flow adjustment and deformation radii, ensuring Ro scaling that translates model results to prototype behaviors in rotating systems. In such experiments, like dam-break currents in , the effective Ro is inferred from the ratio of channel width to Rossby deformation radius (ŵ ≈ 0.25 to 4), where lower ŵ corresponds to smaller Ro and stronger rotational , leading to wall-attached currents and bores that validate nonlinear adjustment theories. In engineering applications, particularly design, the Rossby number assesses Coriolis influences on internal flows, such as in cavities where Ro ≈ 0.4–0.8 at rotational Reynolds numbers Re_θ ≈ 3 × 10^6 – 2 × 10^7 induces flow reversals and unsteady interactions that affect and stability. For example, in rotating disk cavities with axial throughflow, Coriolis effects at these Ro values promote centrifugal buoyancy-driven circulations. Similarly, in low-head hydraulic turbines, Ro values around 7 to 14 quantify Coriolis-induced asymmetries in flows, causing head losses that can be mitigated by adjusting reservoir geometry, as higher latitudes yield lower Ro and greater deflection angles up to 8.3°. Numerical simulations in environments, such as large eddy simulations (LES) of rotating (CFD) models, rely on Ro to guide mesh resolution and capture Coriolis-dominated regimes in flows like turbine blade passages. These approaches validate buoyant flows under these Ro values by resolving unsteady structures and metrics, ensuring simulations align with experimental data on flow stability and phase dynamics. Despite these advances, scaling of the Rossby number faces challenges, particularly in replicating small planetary f values, which often requires elevated velocities and results in hybrid Ro-Re regimes where viscous effects compete with rotation. Dimensional constraints limit experiments to linear scales 10^{-6} times those of natural systems, complicating full replication of and multi-scale processes in low-Ro .

Historical Development

Origin and Naming

The Rossby number emerged in the as part of Carl-Gustaf Rossby's pioneering studies on large-scale , initially conducted at the and later advanced at the following his appointment there in 1940. Rossby's work during this period focused on understanding the dynamics of planetary-scale flows, where rotational effects play a dominant role, laying the groundwork for modern . His seminal 1939 paper, "Relation between variations in the intensity of the zonal circulation of the atmosphere and the displacements of the semi-permanent centers of action," analyzed the propagation and structure of long waves within the westerly winds, highlighting how shapes large-scale atmospheric motions and providing a basis for the scaling parameter that would become known as the Rossby number. The dimensionless number was explicitly formulated in the early through Rossby's research to quantify the relative importance of rotational influences versus inertial forces. In parallel, the same parameter was independently derived by Soviet I. A. Kibel in his 1940 work on perturbations in , leading to its alternative designation as the Kibel number. This contemporaneous recognition underscores the number's foundational role in early , particularly amid rapid progress in theoretical frameworks for during the late 1930s and early .

Key Contributions and Evolution

Following , the Rossby number was integrated into quasi-geostrophic theory, particularly through Jule G. Charney's seminal 1948 paper, which emphasized approximations valid for low Rossby numbers where rotational effects dominate inertial forces in large-scale atmospheric motions. This framework, building on scale analysis, justified the neglect of relative accelerations in the momentum equations for synoptic-scale flows, enabling simplified models of mid-latitude weather systems. In , the concept was extended concurrently by Henry Stommel in his work on wind-driven currents, where the Rossby number helped explain western intensification in subtropical gyres by contrasting interior Sverdrup with narrow layers influenced by the beta effect (the latitudinal variation of the Coriolis parameter). This application highlighted how small Rossby numbers in the interior promote geostrophic , while lateral friction in layers adjusts the . Modern extensions include the local Rossby number, defined as Ro_{local} = \zeta / f, where \zeta is the relative and f is the planetary , used in studies of geophysical to assess the scale-dependent between local rotation and Coriolis effects. This has proven essential in analyzing submesoscale dynamics and vortex instabilities, where Ro_{local} \approx 1 indicates transitional regimes between geostrophic and ageostrophic flows. Influential textbooks have further formalized the Rossby number's role: James R. Holton's 2004 edition of An Introduction to Dynamic Meteorology provides a comprehensive scaling analysis in atmospheric contexts, underscoring its use in deriving balanced equations for weather prediction. Similarly, Lakshmi H. Kantha and Carol A. Clayson's 2000 text Numerical Models of Oceans and Oceanic Processes details its application in oceanic modeling, including parameterization of turbulent closures under varying Rossby regimes. The concept's evolution reflects an initial emphasis on atmospheric dynamics in the 1940s and 1950s, with fuller integration into oceanic circulation theories by the 1960s through advances in frameworks. Today, the Rossby number remains central to modeling, informing the representation of large-scale circulation and wave propagation in coupled atmosphere-ocean general circulation models.

References

  1. [1]
    [PDF] Review of Rossby and topographic waves
    Jun 29, 2023 · ... Rossby number fluid with either continuous stratification (buoyancy frequency = N) or a “1 ½ layer” stratification which has a thin upper ...
  2. [2]
    [PDF] Fluid Dynamics of the Atmosphere and Ocean
    Mar 11, 2018 · The Rossby number characterizes the importance of rotation in a fluid. ... equation, on the basis of small Rossby number. In this chapter we ...
  3. [3]
    [PDF] THREE-DIMENSIONAL QUASI-GEOSTROPHIC SIMULATIONS OF ...
    May 16, 2024 · Swedish – American meteorologist Carl-Gustaf Rossby who was the first to apply fluid dynamics to study the Earth atmosphere: RRoo ...<|control11|><|separator|>
  4. [4]
    [PDF] Equatorial Quasi-Geostrophy - WHOI GFD
    The Rossby number is defined as Ro = U/2ΩL, where U and L are characteristic velocities and lengths, respectively. In the “traditional approximation”, the ...
  5. [5]
    Rossby number
    Rossby number. Non-dimensional ratio (Ro) of ... Ro = U/fL, where U is a characteristic velocity; f, the Coriolis parameter; and L, a characteristic length.
  6. [6]
    10.5 Are all the terms in these equations equally important? Let's ...
    The Coriolis Parameter. Define the Coriolis parameter as f≡2Ωsinϕ. At 45o N, f=(2)(7.27×10−5s−1)sin45o∼10−4s−1 . We will use this parameter in the scale ...
  7. [7]
    [PDF] Geophysical Fluid Dynamics
    Apr 5, 2011 · where the Rossby number Ro is defined as,. Ro = U. f0L . (27). The Rossby number estimates the magnitude of horizontal advection relative to ...Missing: formula | Show results with:formula
  8. [8]
    [PDF] Rotating fluids
    the Rossby number becomes Ro ≈ 104, and the Coriolis force can be completely neglected. The water draining out of your toilet or bathtub moves with similar.
  9. [9]
    [PDF] Elementary Application of the Basic Equations
    The Rossby number is the ratio of the acceleration of the wind to the. Coriolis force. In natural coordinates the Rossby number is given by: When will Ro be ...
  10. [10]
    [PDF] 2 The equations governing atmospheric flow. - Staff
    So, the Rossby number is a dimensionless parameter that comes from scale analysis of momentum balance. ... moisture (see Vallis chapter 1). For, liquids a ...
  11. [11]
    [PDF] ESCI 241 - Meteorology
    The table below summarizes the different regimes of the Rossby number. Regime. Balance. Important Terms. Ro << 1. Geostrophic. Pressure gradient and Coriolis.
  12. [12]
    [PDF] EAPS 52600 Shallow water quasi-geostrophic (QG) system (VallisE ...
    Flow is near-geostrophic balance: small Rossby number, Ro = U. fL ⌧ 1. 2. Horizontal scale of motion similar to deformation radius: L ⇠ Ld (which, with the ...
  13. [13]
    [PDF] What Makes Ocean Gravity Currents Flow Downhill? - WHOI GFD
    Ro = V. fL. (3). Flows with a low ( 1) Rossby number can be considered geostrophic; flows with a Rossby number near 1 must consider the full momentum equations.
  14. [14]
    [PDF] ESCI 344 – Tropical Meteorology
    ... Rossby number (R0 >> 1) implies that the Coriolis term is unimportant, and that the inertial terms and the pressure gradient term balance (cyclostrophic balance) ...Missing: Ro | Show results with:Ro
  15. [15]
    [PDF] 1Steven Brey, 2Nick Davis, 2and Thomas Birner - Colorado State ...
    Mid-latitudes are usually characterized by Rossby numbers ~.1. (consistent with geostrophic balance) where as the tropics are characterized by Rossby ...
  16. [16]
    [PDF] Tropical Cyclone Structure - Meteorology
    Therefore, the core of the storm is in cyclostrophic balance. ○ Outside the core, the Rossby number is of the order of unity, so gradient balance holds.
  17. [17]
    [PDF] 3) Geostrophic wind applies at large space - atmo.arizona.edu
    Assuming that f=10s, we obtain a Rossby number of Ro= V/R ≈ 103, which implies that the Coriolis force can be neglected in computing the balance of forces for ...Missing: cyclones | Show results with:cyclones
  18. [18]
    [PDF] CHAPTER 7 FLOW IN ROTATING ENVIRONMENTS
    The motions of both the atmosphere and the oceans most affected by the. Coriolis acceleration are almost entirely horizontal flows: broad-scale vertical motions ...<|control11|><|separator|>
  19. [19]
    [PDF] 5 Shallow water QG theory. - Staff
    The Rossby number is small but not negligible. It is the small departures from geostrophy that allow the near geostrophic flow of the atmosphere to evolve. The.Missing: dynamics | Show results with:dynamics
  20. [20]
    [PDF] Atmospheric and Oceanic Fluid Dynamics
    Atmospheric and oceanic fluid dynamics (aofd) is both pure and applied. It is a pure because it involves some of the most fundamental and unsolved problems ...
  21. [21]
    [PDF] -36- 5. Rotating Flow The general circulation of the ocean (and ...
    Thus the horizontal Coriolis force must be balanced by the horizontal pressure gradient. This is called geostrophic balance. We hav e already noted that the ...Missing: limitations | Show results with:limitations
  22. [22]
  23. [23]
  24. [24]
    [PDF] Submesoscale Processes and Dynamics - Amala Mahadevan
    This is because flows with. (1) Rossby number exhibit nonlinear Ekman dynamics in which the advection of momentum plays a role in the ekman balance. This ...
  25. [25]
    [PDF] Geostrophic Turbulence
    More specifically, we reqlaire that the Ekman number. (E -= v/f H2), the Rossby number (Ro = U/f L), and the nondimensional frequency [(fT)-1] each be small. U ...
  26. [26]
    [PDF] Fronts and Frontogenesis - University of Wisconsin–Madison
    Extratropical cyclones are characterized by relatively narrow zones of sharp contrast in ... Rossby Number (Ro) is given by Ro=10 m s-1/(10-4 s-1)(106 m ...
  27. [27]
    [PDF] AOSS321_L09_020509_scale_a...
    f = 2 Ω sin(ϕ). Ω ≈ 7.292 × 10-5 s-1. Page 37. Rossby number. • Rossby number: A dimensionless number that indicates the importance of the Coriolis force.
  28. [28]
    [PDF] Quasi-Geostrophic Theory Chapter 5
    Definition: The Rossby number of a flow is a dimensionless quantity which represents the ratio of inertia to Coriolis force. is still used for computing the ...
  29. [29]
    On the violation of gradient wind balance at the top of tropical cyclones
    Aug 12, 2017 · Here Ro = v/fr is the Rossby number, Rog = vg/fr is the geostrophic Rossby number with vg = (g∂Z/∂r)/f as the geostrophic wind (defined to be ...
  30. [30]
    The Golden Radius in Balanced Atmospheric Flows - AMS Journals
    In high-Rossby-number lows (strong winds, small spatial scales, or weak background rotation), gradient balance reduces to cyclostrophic balance. When the Rossby ...
  31. [31]
    [PDF] Waves in the Atmosphere and Oceans - UCI ESS
    Mar 9, 2017 · In atmospheric dynamics and physical oceanography, the Rossby radius of deformation is the length scale at which rotational effects become as.
  32. [32]
    [PDF] Lecture 6: Quasigeostrophic Rossby waves
    The deformation radius is infinite for the barotropic mode (n = 0), and increasingly smaller for each baroclinic mode (n ≥ 1).
  33. [33]
    [PDF] 1.1 Geophysical Fluid Dynamics
    1.2 The Rossby Number 3. (1.2.1). (1.2.2). The nondimensional parameter & is the Rossby number. Large-scale flows are defined as those with sufficiently large ...
  34. [34]
    [PDF] ESCI 342 – Atmospheric Dynamics I
    1 The time scale τ = L /U is called the advective time scale. ... ○ The dimensionless combination U/fL is defined as the Rossby number (named for. Gustav Rossby),.
  35. [35]
    [PDF] Mesoscale Dynamics - twister.ou.edu
    meaningful physically to use the Lagrangian Rossby number Ro, which is defined as the ratio of intrinsic frequency and the Coriolis parameter (. fT f /2. / π.<|control11|><|separator|>
  36. [36]
    [PDF] Geostrophic dynamics at surfaces in the atmosphere and ocean
    small when at small Rossby number Ro ≪ 1, so QG is an accurate approximation. The equation for advection of quasi-geostrophic potential vorticity (QGPV) is.
  37. [37]
    [PDF] Evolution of oceanic circulation theory: From gyres to eddies
    At horizontal scale of 1 km, both the Rossby number and the Richardson number are of order one. Tab. 1 Typical scales for eddies associated with baroclinic ...
  38. [38]
    Geographical Variability of the First Baroclinic Rossby Radius of ...
    Chelton, D. B., and M. G. Schlax, 1996: Global observations of oceanic Rossby waves. Science,272, 234–238. Dewar, W. K., 1998: On “too- ...Introduction · Geographical variability of the... · Discussion and conclusions
  39. [39]
    [PDF] Instability waves in the Gulf Stream front and its thermocline layer
    the separated Gulf Stream the Rossby number is of order one. ... For (Y = 1 and 6 = 1, the constants are: ... (1 + 0~) z + as2 z + o&&2 1 = -C(P, + P,) - C(K ...<|control11|><|separator|>
  40. [40]
    [PDF] EQUATORIAL WAVES - Faculty
    n ¼ 0 indicates the Rossby-gravity (Yanai) wave. Kelvin wave formally corresponds to n ¼ 1. From Gill AE (1982) Atmosphere-Ocean Dynamics, 664pp. New York: ...
  41. [41]
    Global trends in surface eddy mixing from satellite altimetry - Frontiers
    In 73% of the global ocean, changes in eddy mixing length account for more than 50% of the diffusivity trend. The suppressed mixing length theory (SMLT) is ...
  42. [42]
    [PDF] The Effectiveness of Rotating Tank Experiments in Teaching Under
    Feb 16, 2012 · The Rossby number is a dimensionless measure of the ratio of inertial forces to Coriolis forces in a rotating fluid (see. Fig. 3 and discussion ...
  43. [43]
    None
    ### Summary of Rossby Waves in Rotating Annulus Experiments and Rossby Number Role
  44. [44]
    [PDF] Geophysical Fluid Dynamics Laboratory University of Washington
    Dec 19, 2016 · The Rossby number expressing the ratio of fluid e ical o ici o plane a (-) vorticity. • Lab 3: Coriolis force on a micro-jet entering a ...
  45. [45]
    [PDF] Laboratory experiments on nonlinear Rossby adjustment in a channel
    We focus here on the propagation of rotating dam-break gravity currents along a vertical wall. These currents are formed by a finite and instantaneous release ...Missing: number | Show results with:number
  46. [46]
  47. [47]
  48. [48]
  49. [49]
    Carl-Gustaf Rossby: Theorist, institution builder, bon vivant
    Jan 1, 2017 · Rossby joined the Bergen school on 20 June 1919. He arrived as an inquisitive 20-year-old student with no prior knowledge of the atmosphere, and ...
  50. [50]
    [PDF] CARL-GUSTAF ARVID ROSSBY - National Academy of Sciences
    ... 1939, Rossby was studying the large-scale circulations of the atmosphere. This work led to what may be considered his two most important papers (1939,. 1940).
  51. [51]
    [PDF] Carl-Gustaf Rossby - Tellus B
    famous paper of 1939 on the long waves in the westerlies (Rossby, 1939) and the recently pub- lished paper by Staff of the Department of. Meteorology in ...
  52. [52]
    [PDF] Geoffrey K. Vallis - ATMOSPHERIC AND OCEANIC FLUID DYNAMICS
    ... Kibel or Rossby–Kibel number. The notion of geostrophic balance and so, implicitly, that of a small Rossby number, predates both Rossby and Kibel. 10 After ...
  53. [53]
    Numerical Weather Prediction in the Soviet Union - AMS Journals
    I. A. Kibel (1940) derived his simple formula for predicting the motion of ... e is frequently called the "Rossby number" in. American work. A basic ...
  54. [54]
    [PDF] On the scale of atmospheric motions
    Whereas characterized as quasi-hydrostatic, quasi-adiabatic, only the long inertially-propagated waves are quasi-horizontal, and quasi-geostrophic. But here.
  55. [55]
    Rossby Number - an overview | ScienceDirect Topics
    Since the Rossby number is on the order of unity or slightly less, but the Ekman number and the aspect ratio H/L are both much smaller than unity, the Reynolds ...
  56. [56]
    [PDF] the westward intensification of wind-driven ocean currents
    Volume 29, Number 2. April 1948. THE WESTWARD INTENSIFICATION OF WIND-DRIVEN ... lines toward the western border of the ocean. The rest of the streamline ...
  57. [57]
    Generation and Propagation of Inertia–Gravity Waves from Vortex ...
    ... local Rossby number Ro = ζ/f as the ratio between the relative vorticity and planetary vorticity. The local Rossby number turns out to be less applicable ...
  58. [58]
    [PDF] Decaying grid turbulence in a rotating stratified fluid - HAL
    The relative importance of rotation is similarly characterized by the Rossby number: ... A local Rossby number, Rolocal =(ω2 z )1/2/f , can be associated with any ...<|control11|><|separator|>
  59. [59]
    An Introduction to Dynamic Meteorology, Volume 88 - Google Books
    Mar 31, 2004 · An Introduction to Dynamic Meteorology, Fourth Edition presents a cogent explanation of the fundamentals of meteorology, and explains storm dynamics.
  60. [60]
    Resolution issues in numerical models of oceanic and coastal ...
    ... Kantha and Clayson, 2000, Lynch, 2005). For surface gravity wave problems ... In this study, based on the Rossby number similarity criterion, a ...
  61. [61]
    How well is Rossby wave activity represented in the PRIMAVERA ...
    Feb 15, 2022 · This work aims to assess the performance of state-of-the-art global climate models in representing the upper-tropospheric Rossby wave pattern.